| Home | E-Submission | Sitemap | Contact Us |  
top_img
Cancer Research and Treatment > Volume 45(4); 2013 > Article
Phillips, Sheaff, and Szlosarek: Targeting Arginine-Dependent Cancers with Arginine-Degrading Enzymes: Opportunities and Challenges

Abstract

Arginine deprivation is a novel antimetabolite strategy for the treatment of arginine-dependent cancers that exploits differential expression and regulation of key urea cycle enzymes. Several studies have focused on inactivation of argininosuccinate synthetase 1 (ASS1) in a range of malignancies, including melanoma, hepatocellular carcinoma (HCC), mesothelial and urological cancers, sarcomas, and lymphomas. Epigenetic silencing has been identified as a key mechanism for loss of the tumor suppressor role of ASS1 leading to tumoral dependence on exogenous arginine. More recently, dysregulation of argininosuccinate lyase has been documented in a subset of arginine auxotrophic glioblastoma multiforme, HCC and in fumarate hydratase-mutant renal cancers. Clinical trials of several arginine depletors are ongoing, including pegylated arginine deiminase (ADI-PEG20, Polaris Group) and bioengineered forms of human arginase. ADI-PEG20 is furthest along the path of clinical development from combinatorial phase 1 to phase 3 trials and is described in more detail. The challenge will be to identify tumors sensitive to drugs such as ADI-PEG20 and integrate these agents into multimodality drug regimens using imaging and tissue/fluid-based biomarkers as predictors of response. Lastly, resistance pathways to arginine deprivation require further study to optimize arginine-targeted therapies in the oncology clinic.

Introduction

For many human cancers, dysregulation of cellular metabolism is now recognized as a key event in tumor growth and development. Accordingly, research into various aspects of tumor metabolism is undergoing a period of sustained revival, with the hope that this will identify novel tractable pathways for cancer screening, diagnosis and the treatment of malignant disease [1]. Indeed, the development of antimetabolites, such as antifolates and pyrimidine analogues, and the application of Warburg's hypothesis in 18F-fludeoxyglucose positron emission tomography/computed tomography (18F-FDG PET/CT) as a diagnostic and therapeutic imaging tool, demonstrate the importance of studying cancer metabolism in the oncology clinic: continued dissection of the cancer metabolome is likely to further refine our management of patients with cancer. Testing for metabolic genetic drivers such as mutations in Kreb's cycle genes (i.e., fumarate hydratase [FH], succinate dehydrogenase [SDH], isocitrate dehydrogenase-1/2 [IDH 1/2]) is already available, and new PET tracers and antimetabolites are being devised that aim to exploit the differential biochemistry of sugars, lipids and amino acids in normal and malignant cells [2,3]. Here, the focus is on the amino acid L-arginine, which has garnered clinical interest as an exciting cancer target, driven by the emergence of several novel arginine-depleting agents.

Amino Acid Deprivation as an Anticancer Therapeutic Strategy

The rationale for targeting amino acids in cancer therapy was pioneered with the introduction of L-asparaginase in childhood leukemia over five decades ago [4]. In contrast to normal cells, leukemic cells are dependent on exogenous asparagine for growth (i.e., auxotrophic), due to lack of the biosynthetic enzyme asparagine synthetase [5]. As a monotherapy, deamidation of asparagine to aspartate and ammonia by L-asparaginase, led to responses of up to 50% in early studies of patients with chemorefractory leukemia [6]. However, it was not until the advent of multimodality chemotherapy regimens containing L-asparaginase, that the cure rate of acute lymphoblastic leukemia increased from 5% in the 1950's to the present 90%. The problem of hypersensitivity reactions and anti-drug neutralizing antibodies has been addressed to some extent by use of pegylation technology, switching to alternative L-asparaginase enzymes and individualized schedules with close monitoring of plasma asparagine levels [7]. Nevertheless, wider clinical investigation of L-asparaginase, has been hampered by the risk of severe dose-limiting side effects including hepatitis, pancreatitis, coagulopathy and neurotoxicity [8]. Bioengineering efforts are underway to modulate the secondary glutaminolytic activity that mediates, in part, the adverse effects but also the antitumor properties of L-asparaginase [9].
Based on the asparaginase paradigm, several argininecatabolizing enzymes have been developed, including mycoplasma-derived arginine deiminase and human recombinant arginase, for cancer therapy [10]. These enzymes are pegylated for enhanced pharmacokinetics and pharmacodynamics; however, further drug optimization will still be needed for arginine deprivation to gain traction in the clinic [11]. Although the data are preliminary, early signs suggest that arginine deprivation has the potential to make an impact across a broad range of human malignancies.

Arginine Cancer Metabolome: Argininosuccinate Synthetase 1 (ASS1) and Argininosuccinate Lyase (ASL) Enzymes

On an organismal level, diet and protein degradation are the main sources of plasma arginine, with only 5-15% (up to 30% in newborns) derived from de novo biosynthesis via the intestinal-renal axis. Indeed, under physiological conditions the arginine biosynthetic pathway is organized according to tissue function: for example, as the urea cycle in the liver for elimination of nitrogenous waste, and as the nitric oxide (NO)-citrulline cycle in endothelial cells to generate NO, with arginine acting as an intermediary molecule [12].
Numerous studies have confirmed that ASS1, considered a house-keeping gene in normal cells and a rate-limiting biosynthetic enzyme in hepatocytes and endothelial cells, is absent in many tumors [13,14]. Importantly, this leads to a critical dependence on exogenous arginine, with evidence suggesting that tumoral ASS1 deficiency may be both a prognostic biomarker and predictor of sensitivity to arginine deprivation therapy [15]. Methylation-dependent transcriptional silencing has been shown to drive tumoral ASS1 loss in some tumor types, although other mechanisms may also be involved, including repression of the ASS1 promoter by hypoxia-inducible factor-1alpha detected in melanoma cells [16,17]. Further study of the mechanisms of ASS1 loss and sensitivity to arginine depletion will be critical in overcoming resistance in the clinic.
The reason for down-regulation of tumoral ASS1 expression remains unclear and is an area of ongoing research. Recent studies have shown that ASS1 functions as a tumor suppressor in sarcoma and bladder cancer cell lines, reducing colony forming ability, proliferation and invasion of tumor cells and abrogating growth of tumor xenografts in mouse models [18-20]. Indeed, it appears that the reprogramming of tumor arginine metabolism by deactivating ASS1 favors a more aggressive phenotype fueled by exogenous arginine. Previously extracellular arginine has been shown to increase DNA synthesis in Burkitt lymphoma cells, and enhance incorporation of glutamine into nucleotides in colorectal cancer cells [21,22]. Work from our laboratory supports the hypothesis that exogenous arginine is sparing for glutamine in the synthesis of pyrimidines, specifically in ASS1-negative tumor cells. Thus, arginine depletion in ASS1-negative cells led to a marked increase in intracellular glutamine whilst thymidine levels declined secondary to inhibition of thymidylate synthase and dihydrofolate reductase, highlighting the interdependence between exogenous arginine and glutamine for nucleotide synthesis. In contrast, thymidine levels were unaffected by arginine withdrawal in ASS1-positive tumor cells, suggesting that glutamine is being diverted towards arginine synthesis, thereby limiting the capacity for proliferation and growth [20]. Evidence for crosstalk between glutamine and arginine has also been revealed in studies of melanoma cells exposed to ADI-PEG20 [23]. Moreover, metabolomic studies in sarcomas have identified arginine along with glutamine and serine, as essential amino acids with a signature of enhanced gemcitabine sensitivity in a subgroup of liposarcomas [24,25].
Notably, there is also evidence for an aberrant arginine metabolome in nasopharyngeal and bladder cancer cells with an inverse relationship between ASS1 and the expression of key serine pathway enzymes, including phosphoserine phosphatase (PSPH) and phosphoserine aminotransferase I (PSAT) [20,26]. Elevated levels of PSPH and PSAT have been identified in several tumors, and modulate one-carbon folate metabolism [27]. This is an area that will require detailed metabolomic studies to delineate the role of arginine given that there is increasing evidence for differential mechanisms involved in loss of ASS1 expression.
Less is known about the tumoral role of ASL, which is immediately downstream of ASS1 and catalyzes the conversion of argininosuccinate into arginine and fumarate. It is apparent from inborn errors of metabolism that ASL is critical in channeling arginine for NO production, with NO donors reversing the clinical picture of hypertension in children with ASL deficiency [28]. More recently, ASL has been explored in cancer, with upregulation noted in hepatocellular carcinoma (HCC) linked to increased aggressiveness mediated by NO and cyclin A2 signaling [29]. On the other hand, there is evidence that methylated ASL contributes to the arginine auxotrophy of glioblastoma multiforme, with loss of ASS1 and ASL conferring greater sensitivity to ADI-PEG20 [30]. In FH deficient renal cell cancer (RCC), where ASL is intact, the enzyme's activity is reversed producing high levels of argininosuccinate from arginine and fumarate. Here, arginosuccinate levels correlate with reversed ASL activity and may prove to be a biomarker in FH null RCC and other tumors [31,32]. Here too, the role of ASL in human cancer appears to be dependent on tumor type, and further work is needed to understand the implications of ASL expression in human cancer.
Lastly, tumoral heterogeneity of urea-cycle enzyme expression is likely to impact the development of arginine deprivation as a targeted therapy in human cancer. Concerning ASS1, whilst in some tumors, such as melanoma, lymphoma, and small cell lung cancer, there appears to be an 'all or none' level of expression, in the epithelial cancers, non-small cell lung carcinoma, mesothelioma and ovarian cancer, we have noted significant heterogeneity of expression (Fig. 1) [14,16,33-35]. As a result, moving forward into clinical trials we have opted for at least 50% ASS1 negativity within a tumor biopsy for patient selection, thus lowering the bar for patient inclusion, whereas previous trials in melanoma and small cell lung carcinoma (ClinicalTrials.gov Identifier: NCT01266018) have mandated at least 95% ASS1-negative cells.

Arginine Deprivation: Preclinical Data

Several decades of research have led to the notion that arginine promotes tumor growth: manipulation of dietary arginine in cancer cell line studies and in vivo transplantable tumor models; ex vivo analyses of cancer tissue arginase; and in vitro and in vivo pharmacological studies of native arginine-degrading enzymes [36-41]. More recent studies evaluating the efficacy of pegylated bioengineered enzymes have confirmed that arginine depletion is effective in various ASS1-deficient xenograft tumor models, including HCC, melanoma, pancreatic and small cell lung cancer and leukemia [34,42-44]. Whilst amino acid deprivation is known to induce nutritional starvation, the inhibitory effects of arginine depleting agents on tumor growth are not fully understood and remain an area of active investigation. In addition to the links between exogenous arginine and intracellular glutamine and pyrimidine synthesis described above-which impact directly on the cell cycle-ADI-PEG20 also downregulates mammalian target of rapamycin (mTOR) and modulates phosphoinositide 3-kinase (PI3K) via suppression of phosphatase and tensin homolog (PTEN) in ASS1-negative tumor cells [45-49]. It is noteworthy that arginine is one of two key amino acid regulators of mTOR-the other being leucine-and may explain, in part, the ability of arginine depletors to critically impair protein synthesis [50,51]. Apart from modulating tumor biomass via suppressing nucleotide and protein synthesis, arginine also affects tumor cell motility via a NO and focal adhesion kinase (FAK)-dependent pathway; importantly, FAK activation has been linked to deprivation of several other amino acids including glutamine and tyrosine [52-54]. Interestingly, arginine deiminase was identified in an anti-angiogenic screen and an anti-vascular effect has been confirmed both in HUVEC cells in vitro and in a xenograft melanoma study using fluorescence molecular tomography. Furthermore, arginine deiminase was antiangiogenic and potentiated the effect of radiation in a neuroblastoma mouse model [55-58]. Although some native arginine deiminases, found in micro-organisms such as Giardia lamblia, have functions beyond arginine depletion including citrullination of proteins (i.e., a peptidylarginine deiminase), this is not a feature of mycoplasmal-derived ADI-PEG20 used in clinical trials to date [59]. Clearly, arginine-lowering enzymes are likely to impact numerous other pathophysiological pathways and their further characterization will be crucial in optimizing arginine deprivation in the clinic [60-63].
Drug resistance remains a significant challenge for effective usage of arginine depletors. Despite promising results observed in vitro and in vivo as a targeted therapy for arginine auxotrophic tumors, several mechanisms have been identified that mediate tumoral resistance to arginine depletors. The most plausible is re-expression of ASS1, which results in the re-cycling of citrulline back into arginine [64]. This has been described for melanoma cells via transcriptional activation of the ASS1 promoter by the oncogene c-myc [17]. More recently, c-myc has been shown to play a role in the metabolic reprogramming seen during the development of ADI-PEG20 resistance in arginine auxotrophic melanoma cell lines. Here, ADI-PEG20-resistant cells exhibited elevated expression of key enzymes in glutaminolysis, glutaminase (GLS1) and glutamine dehydrogenase, and increased glycolytic metabolic activity, resulting in preferential sensitivity to glutamine inhibitors and glycolytic inhibitors, respectively. c-Myc, not elevated ASS1 expression, was found to contribute to this metabolic rewiring [17,23].
Autophagy ('self-eating'), leading to the provision of a temporary but finite supply of arginine via the breakdown of intracellular organelles, is also triggered by arginine deprivation, circumventing arginine deprivation in several malignancies, including prostate cancer, lymphoma, glioblastoma multiforme, and melanoma [16,30,65,66]. An alternative mechanism may involve macropinocytosis whereby extracellular proteins are degraded to supply the tumor cell with amino acids, as shown recently for K-ras transformed cells deprived of exogenous glutamine [67]. Finally, the direct supply of arginine or its precursors by the tumor microenvironment is also an attractive hypothesis. A similar hypothesis was first proposed and validated for mesenchymal-derived stromal cells providing asparagine for L-asparaginase-resistant acute lymphoblastic leukemia cells and, latterly, adipocytes supplying glutamine, consistent with the dual catalytic activity of L-asparaginase [68,69]. Evidence suggests that tumor associated macrophages (TAMs), predominant cells of the innate immune system within the tumor microenvironment, are involved in complex chemical cross-talk with tumor cells. TAMs are highly versatile and this interplay has the potential to modulate TAM phenotype, which we propose may in turn lead to a form of metabolic cooperation between TAMs and the tumor cells, resulting in the provision of arginine or its precursors to the tumor, thereby by-passing the effect of arginine deprivation. We have recently reviewed the complexity of the tumor microenvironment in relation to arginine, and it is clear that host cells may influence tumorigenesis by being sources of arginase as well as arginine [70-73]. Irrespective of the resistance mechanism, the final common pathway ends with arginine, as this is essential to the survival of the ASS1-negative tumor cell that is unable to supply sufficient amounts of arginine via upregulation of ASS1.
It is likely that further tumoral mechanisms involved in the efficacy and resistance of arginine depletors will be identified and this will be critical in exploiting arginine deprivation with rationally selected drugs for cancer therapy. Ultimately, this may improve the treatment efficacy of many cancers that depend on arginine for survival. To date several preclinical combinatorial studies have shown additive and/or synergistic effects of arginine depletors with taxanes, 5-fluorouracil, pemetrexed cytarabine, PI3K inhibitors, arginine mimetics (e.g., canavanine) and modulators of autophagy using chloroquine and TRAIL [20,44,45,74-77]. Additional interactions that merit further investigation include the potentiation observed between arginine deprivation and radiotherapy in 3D spheroid models of arginine auxotrophic cell lines [78]. Indeed, recent work indicates that ADI-PEG20 is synergistic in vitro and in vivo with radiotherapy in pancreatic cancer cells and may have utility in combination with gemcitabine, a known radiosensitizer and potentiator of ADI-PEG20 in pancreatic malignancy [79,80].
We have investigated further the effect of combining ADI-PEG20 with the antifolate drug pemetrexed based on the novel finding that ADI-PEG20 suppresses the pemetrexed target enzymes, thymidylate synthase (TS) and dihydrofolate reductase (DHFR). It is noteworthy that high TS expression, in particular, has been linked to poor clinical outcome and resistance to antifolates in several cancers including the bladder, lung, mesothelium, and colon [81-85]. Moreover, we showed that loss of ASS1 was correlated with high levels of TS and DHFR and predicted for a poor outcome in patients with transitional carcinoma of the bladder. Subsequently, we validated that arginine deprivation with ADI-PEG20 potentiated the effect of pemetrexed with enhanced apoptosis as assessed by cleaved poly (ADP-ribose) polymerase and annexin V staining. Furthermore, decreased 18F-fluorothymidine (FLT) uptake by PET may be utilized as a predictive imaging biomarker in selecting tumors susceptible to this drug combination (see below).
Specific mention should be made here regarding platinum compounds and their relationship to ASS1 and arginine auxotrophy. Helleman et al. [86] first identified loss of ASS1 as one of 9 genes that predicted independently for cisplatin resistance in a microarray retrospective study of patients with ovarian cancer [87]. Subsequently ASS1 gene overexpression and knock-down in ovarian cancer cells confirmed that while ASS1 loss conferred platinum resistance, there was collateral sensitivity to arginine deprivation [88]. In contrast, studies linking ASS1 loss to enhanced platinum salt sensitivity, namely oxaliplatin and human arginase in HCC cell lines, and cisplatin and ADI-PEG20 in melanoma cells, may exploit a prosurvival pathway that depends on NO synthesis derived from exogenous arginine [89,90]. Consistent with this hypothesis, targeted inhibition of inducible NO synthase synergizes with cisplatin chemotherapy in the treatment of arginine-auxotrophic melanoma cells [91,92].
Lastly, there is increasing interest in the development of novel types of human arginase, in particular a pegylated cobalt-modified (i.e., substituted for native manganese) enzyme (HuArg I [Co]-PEG5K) that has shown good efficacy in depleting arginine [93,94]. Animal toxicity testing indicates that citrulline supplementation is required since arginase, unlike ADI, produces ornithine which cannot be recycled readily to citrulline [95]. Several studies confirm that HuArg I [Co]-PEG5K displays tumor cytotoxicity in vitro and is active in xenograft tumor models [96,97].

Arginine Deprivation: Preclinical Data

The initial phase I/II studies of ADI-PEG20 were performed in patients with liver cancer and melanoma, tumors with a high frequency of ASS1 loss (Table 1) [33,98-105]. Following the first intramuscular injection of 160 IU/m2 of ADI-PEG20, plasma arginine levels decrease by 24 hours and remain low for at least seven days. ADI-PEG20 therapy also triggers a reciprocal increase in plasma citrulline and a decline in plasma nitrite and nitrate levels due to reduced NO synthesis. Response rates of between 25-47% were observed in these small studies with good safety and tolerability [98,99]. The common adverse reactions have been self-limiting grade 1-2 injection site reactions, less frequently diffuse skin rashes and arthralgia, and rarely neutropenia, anaphylactoid reactions and serum sickness. The latest ADI-PEG20 clinical trials have recorded stable disease as the best response, with evidence for a rebound in arginine plasma levels due to drug neutralising antibodies that limit the duration of arginine depletion to a median of 50 days [100,101]. However, as pharmacodymanic measurements were made eight days apart it is possible that more frequent blood testing may have detected oscillations in plasma arginine and this needs further study. Notably, Yang et al. [102] performed a randomized phase 2 study of two different doses of ADI-PEG20 (320 IU/m2 vs. 160 IU/m2) in HCC showing that disease control rates were similar in both groups. A trend towards better survival was observed in patients with >4 weeks (10.0 months) compared with <4 weeks (5.8 months) of arginine deprivation. Based on these encouraging results, a large phase 3 trial (ClinicalTrials.gov Identifier: NCT01 287585) is currently accruing in HCC using the 160 IU/m2 dose with results expected in 2015.
The role of ASS1 as a predictor of response to ADI-PEG20 has been explored only recently in clinical trials. Feun et al. [33] assessed the role of ASS1 expression in a trial of ADI-PEG20 in patients with metastatic melanoma. Responses were confined to patients with ASS1-negative tumors, and were seen only with the higher ADI-PEG20 dose of 320 IU/m2. An improvement in the progression-free survival (PFS) was detected in patients with ASS1-negative compared to ASS1-positive tumors (p=0.025) however no overall survival (OS) difference was observed in this small study. In turn, we have conducted the first prospective multicenter randomized study of ADI-PEG20 in patients pre-selected for ASS1 deficiency in cancer by focusing on the asbestos-related cancer mesothelioma (arginine deiminase and mesothelioma or ADAM). ADAM met its primary endpoint of an improvement in the PFS in the best supportive care (BSC) plus ADI-PEG20 group compared to the BSC alone group of almost 6 weeks (p=0.02) [103]. Our data support differential methylation as the basis for the loss of ASS1 observed in mesothelioma, consistent with our earlier cell line studies.
Thus far, human trials of arginase have been confined to a small phase I study in patients with HCC using peg-rhArgI (BCT, Hong Kong, China), a human pegylated enzyme derived from an Escherichia coli expression system [104]. In contrast to ADI-PEG20, peg-rhArgI was given intravenously (I.V.) and reduced plasma arginine to undetectable levels throughout the treatment period with no evidence for neutralizing antibodies. Interestingly, the PFS and OS for the peg-rhArgI are consistent with that observed in trials of ADI-PEG20 in heavily pre-treated patients with HCC and suggest that the ability to switch on tumoral resistance mechanisms may be critical in determining efficacy to arginine depletors, rather than the duration of arginine depletion per se.

Imaging Arginine Depletion in Human Cancer

Several groups have assessed potential imaging biomarkers of tumor response to arginine deprivation. Ott et al. [101] assessed the effect of arginine deprivation with ADI-PEG20 on metastatic melanoma with both early and late imaging timepoints "using PET." Partial metabolic responses (PMRs) were detected at the early time point in ~25% of patients (day 4), but were not sustained by the late time point (56 days), likely due to a combination of early resistance (e.g., re-expression of ASS1 and induction of autophagy) and a rebound in plasma arginine levels. We also performed early time point imaging (day 15) in the ADAM trial, and while by modified Response Evaluation Criteria in Solid Tumors (RECIST) we confirmed stable disease as the best response, almost 50% of patients had evidence of a PMR by imaging with 18F-FDG PET/CT [103,105].
One of the drawbacks in imaging some arginine auxotrophs such as melanoma, is the apparent increase in glucose uptake following ADI-PEG20 therapy secondary to PTEN loss and PI3K-induced glucose transporter type 1 (GLUT-1) expression [47]. Thus, we and others have explored FLT as an alternative imaging PET tracer and although this does not appear predictive in ASS1-negative melanoma, our studies confirmed robust suppression of FLT uptake following ADI-PEG20 treatment of ASS1-methylated tumors [20,106]. Moreover, the uptake of FLT correlated directly with thymidine-kinase 1 (TK1) protein levels consistent with the mode of action of ADI-PEG20 in modulating protein turnover. A trial is planned to explore the role of FLT in patients with ASS1-deficient thoracic cancers as part of a phase I/II study discussed in more detail below.

Future Directions: Combination Therapies

Targeting arginine is moving towards combinatorial studies based on the emerging preclinical data of additive and synergistic drug interactions in the treatment of arginine auxotrophic cancers. Several ADI-PEG20 combination trials are ongoing or planned and follow the developmental template for asparaginase several decades ago (summarized in Table 2). Studies of the doublet of ADI-PEG20 and cisplatin and ADI-PEG20 and docetaxel to date have shown no untoward toxicity other than that ascribed to either agent alone (Bomalaski J, personal communication).
Building on the ADAM trial, we have hypothesized that ADI-PEG20 with the standard doublet of cisplatin and pemetrexed may provide a novel way of targeting the heterogeneity and chemoresistance of mesothelioma. As previously mentioned, several epithelial tumors, including mesothelioma and non-small cell lung cancer, harbor cells with varying degrees of dependence on exogenous arginine as inferred by the differential expression of ASS1. Since ASS1 positivity is a biomarker of cisplatin sensitivity whereas ASS1 loss is a predictor of cisplatin resistance but collateral sensitivity to ADI-PEG20-pemetrexed, the triplet regimen may be superior to standard chemotherapy alone (Fig. 2). Once the maximum tolerated dose has been defined from the phase 1 triplet combination study, early FLT-PET/CT will be assessed as a potential biomarker of response in patients with thoracic malignancy. Patients will be imaged at baseline and at 24hrs following a single dose of ADI-PEG20 with additional FLTPET/CT scans after the first cycle of chemotherapy and upon completion of all treatment.
From the preclinical studies referred to above, the role of combining arginine deprivation with autophagy inhibitors, glutamine and glycolytic inhibitors, modulators of the tumor microenvironment and radiotherapy, are just a few of the exciting questions to explore in the context of future clinical trials.

Conclusions

Arginine deprivation has the potential to improve the management of several hard-to-treat cancers by exploiting differential expression and regulation of key urea cycle enzymes, ASS1 and ASL. Clinical trials that employ biomarkers for patient selection and treatment response are needed to optimize arginine deprivation therapy in cancer. The field is evolving towards combination strategies and it is hoped that arginine-targeted therapies will soon become part of the standard cancer armamentarium.

Conflicts of Interest

Conflict of interest relevant to this article was not reported.

References

1. Hanahan D, Weinberg RA. Hallmarks of cancer: the next generation. Cell. 2011;144:646–674. PMID: 21376230
crossref pmid
2. Schiffman JD. No child left behind in SDHB testing for paragangliomas and pheochromocytomas. J Clin Oncol. 2011;29:4070–4072. PMID: 21969491
crossref pmid
3. Vander Heiden MG. Targeting cancer metabolism: a therapeutic window opens. Nat Rev Drug Discov. 2011;10:671–684. PMID: 21878982
crossref pmid
4. Broome JD. Evidence that the L-asparaginase activity of Guinea pig serum is responsible for its antilymphoma effects. Nature. 1961;191:1114–1115.
crossref
5. Richards NG, Kilberg MS. Asparagine synthetase chemotherapy. Annu Rev Biochem. 2006;75:629–654. PMID: 16756505
crossref pmid pmc
6. Jaffe N, Traggis D, Das L, Frauenberger G, Hann HW, Kim BS, et al. Favorable remission induction rate with twice weekly doses of L-asparaginase. Cancer Res. 1973;33:1–4. PMID: 4565907
pmid
7. Douer D. Is asparaginase a critical component in the treatment of acute lymphoblastic leukemia? Best Pract Res Clin Haematol. 2008;21:647–658. PMID: 19041604
crossref pmid
8. Truelove E, Fielding AK, Hunt BJ. The coagulopathy and thrombotic risk associated with L-asparaginase treatment in adults with acute lymphoblastic leukaemia. Leukemia. 2013;27:553–559. PMID: 23099335
crossref pmid
9. Offman MN, Krol M, Patel N, Krishnan S, Liu J, Saha V, et al. Rational engineering of L-asparaginase reveals importance of dual activity for cancer cell toxicity. Blood. 2011;117:1614–1621. PMID: 21106986
crossref pmid
10. Wheatley DN. Controlling cancer by restricting arginine availability: arginine-catabolizing enzymes as anticancer agents. Anticancer Drugs. 2004;15:825–833. PMID: 15457122
crossref pmid
11. Feun L, Savaraj N. Pegylated arginine deiminase: a novel anticancer enzyme agent. Expert Opin Investig Drugs. 2006;15:815–822.
crossref pmid pmc
12. Husson A, Brasse-Lagnel C, Fairand A, Renouf S, Lavoinne A. Argininosuccinate synthetase from the urea cycle to the citrulline-NO cycle. Eur J Biochem. 2003;270:1887–1899. PMID: 12709047
crossref pmid
13. Dillon BJ, Prieto VG, Curley SA, Ensor CM, Holtsberg FW, Bomalaski JS, et al. Incidence and distribution of argininosuccinate synthetase deficiency in human cancers: a method for identifying cancers sensitive to arginine deprivation. Cancer. 2004;100:826–833. PMID: 14770441
crossref pmid
14. Szlosarek PW, Grimshaw MJ, Wilbanks GD, Hagemann T, Wilson JL, Burke F, et al. Aberrant regulation of argininosuccinate synthetase by TNF-alpha in human epithelial ovarian cancer. Int J Cancer. 2007;121:6–11. PMID: 17354225
crossref pmid
15. Delage B, Fennell DA, Nicholson L, McNeish I, Lemoine NR, Crook T, et al. Arginine deprivation and argininosuccinate synthetase expression in the treatment of cancer. Int J Cancer. 2010;126:2762–2772. PMID: 20104527
crossref pmid
16. Delage B, Luong P, Maharaj L, O'Riain C, Syed N, Crook T, et al. Promoter methylation of argininosuccinate synthetase-1 sensitises lymphomas to arginine deiminase treatment, autophagy and caspase-dependent apoptosis. Cell Death Dis. 2012;3:e342PMID: 22764101
crossref pmid pmc
17. Tsai WB, Aiba I, Lee SY, Feun L, Savaraj N, Kuo MT. Resistance to arginine deiminase treatment in melanoma cells is associated with induced argininosuccinate synthetase expression involving c-Myc/HIF-1alpha/Sp4. Mol Cancer Ther. 2009;8:3223–3233. PMID: 19934275
crossref pmid pmc
18. Kobayashi E, Masuda M, Nakayama R, Ichikawa H, Satow R, Shitashige M, et al. Reduced argininosuccinate synthetase is a predictive biomarker for the development of pulmonary metastasis in patients with osteosarcoma. Mol Cancer Ther. 2010;9:535–544. PMID: 20159990
crossref pmid
19. Huang HY, Wu WR, Wang YH, Wang JW, Fang FM, Tsai JW, et al. ASS1 as a novel tumor suppressor gene in myxofibrosarcomas: aberrant loss via epigenetic DNA methylation confers aggressive phenotypes, negative prognostic impact, and therapeutic relevance. Clin Cancer Res. 2013;19:2861–2872. PMID: 23549872
crossref pmid
20. Allen M, Luong P, Hudson C, Leyton J, Delage B, Ghazaly E, et al. Prognostic and therapeutic impact of argininosuccinate synthetase-1 control in bladder cancer as monitored longitudinally by PET imaging. Cancer Res. 2013;11. 27[Epub]. http://dx.doi.org/10.1158/0008-5472.CAN-13-1702
crossref
21. Osunkoya BO, Adler WH, Smith RT. Effect of arginine deficiency on synthesis of DNA and immunoglobulin receptor of Burkitt lymphoma cells. Nature. 1970;227:398–399. PMID: 4193643
crossref pmid
22. Yamauchi K, Komatsu T, Kulkarni AD, Ohmori Y, Minami H, Ushiyama Y, et al. Glutamine and arginine affect Caco-2 cell proliferation by promotion of nucleotide synthesis. Nutrition. 2002;18:329–333. PMID: 11934546
crossref pmid
23. Long Y, Tsai WB, Wangpaichitr M, Tsukamoto T, Savaraj N, Feun LG, et al. Arginine deiminase resistance in melanoma cells is associated with metabolic reprogramming, glucose dependence, and glutamine addiction. Mol Cancer Ther. 2013;12:2581–2590. PMID: 23979920
crossref pmid pmc
24. Braas D, Ahler E, Tam B, Nathanson D, Riedinger M, Benz MR, et al. Metabolomics strategy reveals subpopulation of liposarcomas sensitive to gemcitabine treatment. Cancer Discov. 2012;2:1109–1117. PMID: 23230188
crossref pmid pmc
25. Van Tine BA, Bean GR, Boone P, Tanas M, Schulze MB, Chen DY, et al. Using pegylated arginine deiminase (ADI-PEG20) for the treatment of sarcomas that lack argininosuccinate synthesase 1 expression. J Clin Oncol. 2013;31(S):10526.
crossref
26. Lan J, Tai HC, Lee SW, Chen TJ, Huang HY, Li CF. Deficiency in expression and epigenetic DNA methylation of ASS1 gene in nasopharyngeal carcinoma: negative prognostic impact and therapeutic relevance. Tumour Biol. 2013;7. 30[Epub]. http://dx.doi.org/10.1007/s13277-013-1020-8
crossref
27. Possemato R, Marks KM, Shaul YD, Pacold ME, Kim D, Birsoy K, et al. Functional genomics reveal that the serine synthesis pathway is essential in breast cancer. Nature. 2011;476:346–350. PMID: 21760589
crossref pmid pmc
28. Erez A, Nagamani SC, Shchelochkov OA, Premkumar MH, Campeau PM, Chen Y, et al. Requirement of argininosuccinate lyase for systemic nitric oxide production. Nat Med. 2011;17:1619–1626. PMID: 22081021
crossref pmid pmc
29. Huang HL, Hsu HP, Shieh SC, Chang YS, Chen WC, Cho CY, et al. Attenuation of argininosuccinate lyase inhibits cancer growth via cyclin A2 and nitric oxide. Mol Cancer Ther. 2013;12:2505–2516. PMID: 23979921
crossref pmid
30. Syed N, Langer J, Janczar K, Singh P, Lo Nigro C, Lattanzio L, et al. Epigenetic status of argininosuccinate synthetase and argininosuccinate lyase modulates autophagy and cell death in glioblastoma. Cell Death Dis. 2013;4:e458PMID: 23328665
crossref pmid pmc
31. Zheng L, MacKenzie ED, Karim SA, Hedley A, Blyth K, Kalna G, et al. Reversed argininosuccinate lyase activity in fumarate hydratase-deficient cancer cells. Cancer Metab. 2013;1:12PMID: 24280230
crossref pmid pmc
32. Adam J, Yang M, Bauerschmidt C, Kitagawa M, O'Flaherty L, Maheswaran P, et al. A role for cytosolic fumarate hydratase in urea cycle metabolism and renal neoplasia. Cell Rep. 2013;3:1440–1448. PMID: 23643539
crossref pmid pmc
33. Feun LG, Marini A, Walker G, Elgart G, Moffat F, Rodgers SE, et al. Negative argininosuccinate synthetase expression in melanoma tumours may predict clinical benefit from arginine-depleting therapy with pegylated arginine deiminase. Br J Cancer. 2012;106:1481–1485. PMID: 22472884
crossref pmid pmc
34. Kelly MP, Jungbluth AA, Wu BW, Bomalaski J, Old LJ, Ritter G. Arginine deiminase PEG20 inhibits growth of small cell lung cancers lacking expression of argininosuccinate synthetase. Br J Cancer. 2012;106:324–332. PMID: 22134507
crossref pmid pmc
35. Szlosarek PW, Klabatsa A, Pallaska A, Sheaff M, Smith P, Crook T, et al. In vivo loss of expression of argininosuccinate synthetase in malignant pleural mesothelioma is a biomarker for susceptibility to arginine depletion. Clin Cancer Res. 2006;12:7126–7131. PMID: 17145837
crossref pmid
36. Gilroy E. The influence of arginine upon the growth rate of a transplantable tumour in the mouse. Biochem J. 1930;24:589–595. PMID: 16744397
crossref pmid pmc
37. Yeatman TJ, Risley GL, Brunson ME. Depletion of dietary arginine inhibits growth of metastatic tumor. Arch Surg. 1991;126:1376–1381. PMID: 1747050
crossref pmid
38. Bach SJ, Lasnitzki I. Some aspects of the role of arginine and arginase in mouse carcinoma 63. Enzymologia. 1947;12:198–205. PMID: 18910560
pmid
39. Storr JM, Burton AF. The effects of arginine deficiency on lymphoma cells. Br J Cancer. 1974;30:50–59. PMID: 4528778
crossref pmid pmc
40. Scott L, Lamb J, Smith S, Wheatley DN. Single amino acid (arginine) deprivation: rapid and selective death of cultured transformed and malignant cells. Br J Cancer. 2000;83:800–810. PMID: 10952786
crossref pmid pmc
41. Takaku H, Takase M, Abe S, Hayashi H, Miyazaki K. In vivo anti-tumor activity of arginine deiminase purified from Mycoplasma arginini. Int J Cancer. 1992;51:244–249. PMID: 1568792
crossref pmid
42. Ensor CM, Holtsberg FW, Bomalaski JS, Clark MA. Pegylated arginine deiminase (ADI-SS PEG20,000 mw) inhibits human melanomas and hepatocellular carcinomas in vitro and in vivo. Cancer Res. 2002;62:5443–5450. PMID: 12359751
pmid
43. Bowles TL, Kim R, Galante J, Parsons CM, Virudachalam S, Kung HJ, et al. Pancreatic cancer cell lines deficient in argininosuccinate synthetase are sensitive to arginine deprivation by arginine deiminase. Int J Cancer. 2008;123:1950–1955. PMID: 18661517
crossref pmid pmc
44. Hernandez CP, Morrow K, Lopez-Barcons LA, Zabaleta J, Sierra R, Velasco C, et al. Pegylated arginase I: a potential therapeutic approach in T-ALL. Blood. 2010;115:5214–5221. PMID: 20407034
crossref pmid pmc
45. Kim RH, Coates JM, Bowles TL, McNerney GP, Sutcliffe J, Jung JU, et al. Arginine deiminase as a novel therapy for prostate cancer induces autophagy and caspase-independent apoptosis. Cancer Res. 2009;69:700–708. PMID: 19147587
crossref pmid pmc
46. Wu FL, Liang YF, Chang YC, Yo HH, Wei MF, Shen LJ. RNA interference of argininosuccinate synthetase restores sensitivity to recombinant arginine deiminase (rADI) in resistant cancer cells. J Biomed Sci. 2011;18:25PMID: 21453546
crossref pmid pmc
47. Stelter L, Evans MJ, Jungbluth AA, Zanzonico P, Ritter G, Ku T, et al. Novel mechanistic insights into arginine deiminase pharmacology suggest 18F-FDG is not suitable to evaluate clinical response in melanoma. J Nucl Med. 2012;53:281–286. PMID: 22228793
crossref pmid
48. Gong H, Zolzer F, von Recklinghausen G, Rossler J, Breit S, Havers W, et al. Arginine deiminase inhibits cell proliferation by arresting cell cycle and inducing apoptosis. Biochem Biophys Res Commun. 1999;261:10–14. PMID: 10405315
crossref pmid
49. Gong H, Zolzer F, von Recklinghausen G, Havers W, Schweigerer L. Arginine deiminase inhibits proliferation of human leukemia cells more potently than asparaginase by inducing cell cycle arrest and apoptosis. Leukemia. 2000;14:826–829. PMID: 10803513
crossref pmid
50. Morrow K, Hernandez CP, Raber P, Del Valle L, Wilk AM, Majumdar S, et al. Anti-leukemic mechanisms of pegylated arginase I in acute lymphoblastic T-cell leukemia. Leukemia. 2013;27:569–577. PMID: 22926702
crossref pmid pmc
51. Laplante M, Sabatini DM. mTOR signaling in growth control and disease. Cell. 2012;149:274–293. PMID: 22500797
crossref pmid pmc
52. Rhoads JM, Chen W, Gookin J, Wu GY, Fu Q, Blikslager AT, et al. Arginine stimulates intestinal cell migration through a focal adhesion kinase dependent mechanism. Gut. 2004;53:514–522. PMID: 15016745
crossref pmid pmc
53. Rhoads JM, Liu Y, Niu X, Surendran S, Wu G. Arginine stimulates cdx2-transformed intestinal epithelial cell migration via a mechanism requiring both nitric oxide and phosphorylation of p70 S6 kinase. J Nutr. 2008;138:1652–1657. PMID: 18716165
crossref pmid
54. Fu YM, Zhang H, Ding M, Li YQ, Fu X, Yu ZX, et al. Specific amino acid restriction inhibits attachment and spreading of human melanoma via modulation of the integrin/focal adhesion kinase pathway and actin cytoskeleton remodeling. Clin Exp Metastasis. 2004;21:587–598. PMID: 15787096
crossref pmid
55. Wilm M, Shevchenko A, Houthaeve T, Breit S, Schweigerer L, Fotsis T, et al. Femtomole sequencing of proteins from polyacrylamide gels by nano-electrospray mass spectrometry. Nature. 1996;379:466–469. PMID: 8559255
crossref pmid
56. Park IS, Kang SW, Shin YJ, Chae KY, Park MO, Kim MY, et al. Arginine deiminase: a potential inhibitor of angiogenesis and tumour growth. Br J Cancer. 2003;89:907–914. PMID: 12942125
crossref pmid pmc
57. Stelter L, Evans MJ, Jungbluth AA, Longo VA, Zanzonico P, Ritter G, et al. Imaging of tumor vascularization using fluorescence molecular tomography to monitor arginine deiminase treatment in melanoma. Mol Imaging. 2013;12:67–73. PMID: 23348793
pmid
58. Gong H, Pottgen C, Stuben G, Havers W, Stuschke M, Schweigerer L. Arginine deiminase and other antiangiogenic agents inhibit unfavorable neuroblastoma growth: potentiation by irradiation. Int J Cancer. 2003;106:723–728. PMID: 12866032
crossref pmid
59. Touz MC, Ropolo AS, Rivero MR, Vranych CV, Conrad JT, Svard SG, et al. Arginine deiminase has multiple regulatory roles in the biology of Giardia lamblia. J Cell Sci. 2008;121(Pt 17):2930–2938. PMID: 18697833
crossref pmid pmc
60. Morris SM Jr. Recent advances in arginine metabolism: roles and regulation of the arginases. Br J Pharmacol. 2009;157:922–930. PMID: 19508396
crossref pmid pmc
61. Wu G, Morris SM Jr. Arginine metabolism: nitric oxide and beyond. Biochem J. 1998;336(Pt 1):1–17. PMID: 9806879
crossref pmid pmc
62. Leong HX, Simkevich C, Lesieur-Brooks A, Lau BW, Fugere C, Sabo E, et al. Short-term arginine deprivation results in large-scale modulation of hepatic gene expression in both normal and tumor cells: microarray bioinformatic analysis. Nutr Metab (Lond). 2006;3:37PMID: 16961918
crossref pmid pmc
63. Brasse-Lagnel CG, Lavoinne AM, Husson AS. Amino acid regulation of mammalian gene expression in the intestine. Biochimie. 2010;92:729–735. PMID: 20188788
crossref pmid
64. Shen LJ, Lin WC, Beloussow K, Shen WC. Resistance to the anti-proliferative activity of recombinant arginine deiminase in cell culture correlates with the endogenous enzyme, argininosuccinate synthetase. Cancer Lett. 2003;191:165–170. PMID: 12618329
crossref pmid
65. Kim RH, Bold RJ, Kung HJ. ADI, autophagy and apoptosis: metabolic stress as a therapeutic option for prostate cancer. Autophagy. 2009;5:567–568. PMID: 19276647
crossref pmid pmc
66. Wang Z, Shi X, Li Y, Zeng X, Fan J, Sun Y, et al. Involvement of autophagy in recombinant human arginase-induced cell apoptosis and growth inhibition of malignant melanoma cells. Appl Microbiol Biotechnol. 2013;8. 06[Epub]. http://dx.doi.org/10.1007/s00253-013-5118-0
crossref
67. Commisso C, Davidson SM, Soydaner-Azeloglu RG, Parker SJ, Kamphorst JJ, Hackett S, et al. Macropinocytosis of protein is an amino acid supply route in Ras-transformed cells. Nature. 2013;497:633–637. PMID: 23665962
crossref pmid pmc
68. Iwamoto S, Mihara K, Downing JR, Pui CH, Campana D. Mesenchymal cells regulate the response of acute lymphoblastic leukemia cells to asparaginase. J Clin Invest. 2007;117:1049–1057. PMID: 17380207
crossref pmid pmc
69. Ehsanipour EA, Sheng X, Behan JW, Wang X, Butturini A, Avramis VI, et al. Adipocytes cause leukemia cell resistance to L-asparaginase via release of glutamine. Cancer Res. 2013;73:2998–3006. PMID: 23585457
crossref pmid pmc
70. Ellyard JI, Quah BJ, Simson L, Parish CR. Alternatively activated macrophage possess antitumor cytotoxicity that is induced by IL-4 and mediated by arginase-1. J Immunother. 2010;33:443–452. PMID: 20463604
crossref pmid
71. Zea AH, Rodriguez PC, Atkins MB, Hernandez C, Signoretti S, Zabaleta J, et al. Arginase-producing myeloid suppressor cells in renal cell carcinoma patients: a mechanism of tumor evasion. Cancer Res. 2005;65:3044–3048. PMID: 15833831
crossref pmid
72. Kwong-Lam F, Chi-Fung CG. Vincristine could partly suppress stromal support to T-ALL blasts during pegylated arginase I treatment. Exp Hematol Oncol. 2013;2:11PMID: 23574711
crossref pmid pmc
73. Phillips M, Szlosarek PW. Arginine metabolism and tumour-associated macrophagesIn : Lawrence T, Hagemann T, editors. Tumour-associated macrophages. New York: Springer; 2012. p. 77–90.

74. Cheng PN, Lam TL, Lam WM, Tsui SM, Cheng AW, Lo WH, et al. Pegylated recombinant human arginase (rhArgpeg5,000mw) inhibits the in vitro and in vivo proliferation of human hepatocellular carcinoma through arginine depletion. Cancer Res. 2007;67:309–317. PMID: 17210712
crossref pmid
75. Tsai WB, Aiba I, Long Y, Lin HK, Feun L, Savaraj N, et al. Activation of Ras/PI3K/ERK pathway induces c-Myc stabilization to upregulate argininosuccinate synthetase, leading to arginine deiminase resistance in melanoma cells. Cancer Res. 2012;72:2622–2633. PMID: 22461507
crossref pmid pmc
76. You M, Savaraj N, Kuo MT, Wangpaichitr M, Varona-Santos J, Wu C, et al. TRAIL induces autophagic protein cleavage through caspase activation in melanoma cell lines under arginine deprivation. Mol Cell Biochem. 2013;374:181–190. PMID: 23180246
crossref pmid pmc
77. Vynnytska BO, Mayevska OM, Kurlishchuk YV, Bobak YP, Stasyk OV. Canavanine augments proapoptotic effects of arginine deprivation in cultured human cancer cells. Anticancer Drugs. 2011;22:148–157. PMID: 20717004
crossref pmid
78. Vynnytska-Myronovska B, Bobak Y, Garbe Y, Dittfeld C, Stasyk O, Kunz-Schughart LA. Single amino acid arginine starvation efficiently sensitizes cancer cells to canavanine treatment and irradiation. Int J Cancer. 2012;130:2164–2175. PMID: 21647872
crossref pmid
80. Deorukhkar A, Diep N, Chatterjee D, Diagardjane P, Bomalaski J, Krishnan S. Arginine deiminase: a novel radiosensitizer in pancreatic cancer in vitro and in vivo. In: 2014 ASCO Gastrointestinal Cancers Symposium; 2014 Jan 16-18; San Francisco, CA. Abstr no. 221

80. Daylami R, Muilenburg D, Bowles TL, Martinez SR, Bold RJ. Arginine deprivation by PEG-ADI induces autophagic cell death and enhances the tumor suppression effect of gemcitabine in pancreatic cancer. In: 2010 AACR 101st Annual Meeting; 2010 April 17-21; Washington, DC. Abstr no. 484

81. Li Y, Li X, Dai H, Sun X, Li J, Yang F, et al. Thymidylate synthase was associated with patient prognosis and the response to adjuvant therapy in bladder cancer. BJU Int. 2009;103:547–552. PMID: 18990150
crossref pmid
82. Takezawa K, Okamoto I, Okamoto W, Takeda M, Sakai K, Tsukioka S, et al. Thymidylate synthase as a determinant of pemetrexed sensitivity in non-small cell lung cancer. Br J Cancer. 2011;104:1594–1601. PMID: 21487406
crossref pmid pmc
83. Righi L, Papotti MG, Ceppi P, Bille A, Bacillo E, Molinaro L, et al. Thymidylate synthase but not excision repair crosscomplementation group 1 tumor expression predicts outcome in patients with malignant pleural mesothelioma treated with pemetrexed-based chemotherapy. J Clin Oncol. 2010;28:1534–1539. PMID: 20177021
crossref pmid
84. Sigmond J, Backus HH, Wouters D, Temmink OH, Jansen G, Peters GJ. Induction of resistance to the multitargeted antifolate Pemetrexed (ALIMTA) in WiDr human colon cancer cells is associated with thymidylate synthase overexpression. Biochem Pharmacol. 2003;66:431–438. PMID: 12907242
crossref pmid
85. Edler D, Kressner U, Ragnhammar P, Johnston PG, Magnusson I, Glimelius B, et al. Immunohistochemically detected thymidylate synthase in colorectal cancer: an independent prognostic factor of survival. Clin Cancer Res. 2000;6:488–492. PMID: 10690528
pmid
86. Helleman J, Jansen MP, Span PN, van Staveren IL, Massuger LF, Meijer-van Gelder ME, et al. Molecular profiling of platinum resistant ovarian cancer. Int J Cancer. 2006;118:1963–1971. PMID: 16287073
crossref pmid
87. Melaiu O, Cristaudo A, Melissari E, Di Russo M, Bonotti A, Bruno R, et al. A review of transcriptome studies combined with data mining reveals novel potential markers of malignant pleural mesothelioma. Mutat Res. 2012;750:132–140. PMID: 22198210
crossref pmid
88. Nicholson LJ, Smith PR, Hiller L, Szlosarek PW, Kimberley C, Sehouli J, et al. Epigenetic silencing of argininosuccinate synthetase confers resistance to platinum-induced cell death but collateral sensitivity to arginine auxotrophy in ovarian cancer. Int J Cancer. 2009;125:1454–1463. PMID: 19533750
crossref pmid
89. Chow AK, Ng L, Sing Li H, Cheng CW, Lam CS, Yau TC, et al. Anti-tumor efficacy of a recombinant human arginase in human hepatocellular carcinoma. Curr Cancer Drug Targets. 2012;12:1233–1243. PMID: 22873218
crossref pmid
90. Savaraj N, You M, Wu C, Wangpaichitr M, Kuo MT, Feun LG. Arginine deprivation, autophagy, apoptosis (AAA) for the treatment of melanoma. Curr Mol Med. 2010;10:405–412. PMID: 20459375
crossref pmid pmc
91. Sikora AG, Gelbard A, Davies MA, Sano D, Ekmekcioglu S, Kwon J, et al. Targeted inhibition of inducible nitric oxide synthase inhibits growth of human melanoma in vivo and synergizes with chemotherapy. Clin Cancer Res. 2010;16:1834–1844. PMID: 20215556
crossref pmid pmc
92. Grimm EA, Sikora AG, Ekmekcioglu S. Molecular pathways: inflammation-associated nitric-oxide production as a cancersupporting redox mechanism and a potential therapeutic target. Clin Cancer Res. 2013;19:5557–5563. PMID: 23868870
crossref pmid pmc
93. Stone E, Chantranupong L, Gonzalez C, O'Neal J, Rani M, VanDenBerg C, et al. Strategies for optimizing the serum persistence of engineered human arginase I for cancer therapy. J Control Release. 2012;158:171–179. PMID: 22001609
crossref pmid pmc
94. Agrawal V, Woo JH, Mauldin JP, Jo C, Stone EM, Georgiou G, et al. Cytotoxicity of human recombinant arginase I (Co)-PEG5000 in the presence of supplemental L-citrulline is dependent on decreased argininosuccinate synthetase expression in human cells. Anticancer Drugs. 2012;23:51–64. PMID: 21955999
crossref pmid
95. Mauldin JP, Zeinali I, Kleypas K, Woo JH, Blackwood RS, Jo CH, et al. Recombinant human arginase toxicity in mice is reduced by citrulline supplementation. Transl Oncol. 2012;5:26–31. PMID: 22348173
crossref pmid pmc
96. Glazer ES, Stone EM, Zhu C, Massey KL, Hamir AN, Curley SA. Bioengineered human arginase I with enhanced activity and stability controls hepatocellular and pancreatic carcinoma xenografts. Transl Oncol. 2011;4:138–146. PMID: 21633669
crossref pmid pmc
97. Tanios R, Bekdash A, Kassab E, Stone E, Georgiou G, Frankel AE, et al. Human recombinant arginase I(Co)-PEG5000 [HuArgI(Co)-PEG5000]-induced arginine depletion is selectively cytotoxic to human acute myeloid leukemia cells. Leuk Res. 2013;37:1565–1571. PMID: 24018014
crossref pmid
98. Izzo F, Marra P, Beneduce G, Castello G, Vallone P, De Rosa V, et al. Pegylated arginine deiminase treatment of patients with unresectable hepatocellular carcinoma: results from phase I/II studies. J Clin Oncol. 2004;22:1815–1822. PMID: 15143074
crossref pmid
99. Ascierto PA, Scala S, Castello G, Daponte A, Simeone E, Ottaiano A, et al. Pegylated arginine deiminase treatment of patients with metastatic melanoma: results from phase I and II studies. J Clin Oncol. 2005;23:7660–7668. PMID: 16234528
crossref pmid
100. Glazer ES, Piccirillo M, Albino V, Di Giacomo R, Palaia R, Mastro AA, et al. Phase II study of pegylated arginine deiminase for nonresectable and metastatic hepatocellular carcinoma. J Clin Oncol. 2010;28:2220–2226. PMID: 20351325
crossref pmid
101. Ott PA, Carvajal RD, Pandit-Taskar N, Jungbluth AA, Hoffman EW, Wu BW, et al. Phase I/II study of pegylated arginine deiminase (ADI-PEG 20) in patients with advanced melanoma. Invest New Drugs. 2013;31:425–434. PMID: 22864522
crossref pmid pmc
102. Yang TS, Lu SN, Chao Y, Sheen IS, Lin CC, Wang TE, et al. A randomised phase II study of pegylated arginine deiminase (ADI-PEG 20) in Asian advanced hepatocellular carcinoma patients. Br J Cancer. 2010;103:954–960. PMID: 20808309
crossref pmid pmc
103. Szlosarek PW, Steele JP, Sheaff MT, Avril NE, Szyszko T, Ellis S, et al. A randomised phase II trial of pegylated arginine deiminase (ADI-PEG20) in patients with malignant pleural mesothelioma (MPM). In: 2013 World Conference on Lung Cancer; 2013 Oct 27-30; Sydney. Abstr no. MO09.02

104. Yau T, Cheng PN, Chan P, Chan W, Chen L, Yuen J, et al. A phase 1 dose-escalating study of pegylated recombinant human arginase 1 (Peg-rhArg1) in patients with advanced hepatocellular carcinoma. Invest New Drugs. 2013;31:99–107. PMID: 22426640
crossref pmid pmc
105. Szlosarek PW, Luong P, Phillips MM, Baccarini M, Stephen E, Szyszko T, et al. Metabolic response to pegylated arginine deiminase in mesothelioma with promoter methylation of argininosuccinate synthetase. J Clin Oncol. 2013;31:e111–e113. PMID: 23319692
crossref pmid
106. Stelter L, Fuchs S, Jungbluth AA, Ritter G, Longo VA, Zanzonico P, et al. Evaluation of arginine deiminase treatment in melanoma xenografts using (18)F-FLT PET. Mol Imaging Biol. 2013;15:768–775. PMID: 23722880
crossref pmid pmc
Fig. 1
Heterogeneity of argininosuccinate synthetase 1 (ASS1) expression in lung adenocarcinoma. Immunohistochemical staining (BD ASS1 antibody 1:500, BioGenex Super Sensitive Detection System) of a non-small cell lung cancer composed of a mixed population of ASS1 positive and negative tumor cells (×400).
crt-45-251-g001.jpg
Fig. 2
Multiagent chemotherapy for arginine auxotrophic cancers. Applying the analogy of ten-pin bowling, a 'strike' is rarely seen in cancer therapy whereas a 'half-strike' is more likely with multiagent drug regimens; various drug combination studies are underway including the example shown here of the triplet, ADI-PEG20, cisplatin and pemetrexed. ASS1, argininosuccinate synthetase 1.
crt-45-251-g002.jpg
Table 1
Completed clinical studies of single-agent arginine depletors in advanced cancer
Study No. of patients Tumor type Arginine depletor Overall survival (mo) Response rate (best response) Comments
Izzo et al. (2004) [98] 19 HCC ADI-PEG20 phase I/II 13.7 47% (CR+PR) Low toxicity
Glazer et al. (2010) [100] 76 HCC ADI-PEG20 phase II 15.8a) SD only [Arginine] plasma reduced
Yang et al. (2010) [102] 71 HCC Randomized phase II: 160 or 320 IU/m2 of ADI-PEG20 7.3 31% (SD) PFS 1.8 mo; heavily pre-treated patients
Yau et al.(2013) [104] 15 HCC peg-rhArg1 phase I 5.1 26.7% (SD) >8 wk PFS 2.8 mo; heavily pre-treated; OBD=1,600 IU/kg; I.V. administration
Ascierto et al. (2005) [99] 39 Melanoma ADI-PEG20 phase I/II 15 25% (CR+PR) Reduced NO synthesis consistent with MOA
Feun et al. (2012) [33] 38 Melanoma ADI-PEG20 phase II 14.6 (ASS1-ve) vs.
9.3 (ASS1+ve) (p=0.374)
11% (PR) in ASS1-ve only PFS 3.6 mo (ASS1-ve) vs. 1.8 mo (ASS1+ve) (p=0.025); 74% ASS1-ve frequency; ASS1 re-expression on relapse (n=2/2)
Ott et al. (2013) [101] 31 Melanoma ADI-PEG20 phase I/II n/a 25% Early PMR, 29% (SD) Prolonged SD in choroidal melanoma
Szlosarek et al. (2013) [103,105] 68 Mesothelioma Randomized phase II in ASS1-ve patients: ADI- PEG20+BSC (n=44) vs. BSC alone (n=24) n/a 46% Early PMR, SD by modified RECIST PFS 3.3 mo vs. 1.9 mo (p=0.02) favoring ADI-PEG20+BSC; ASS1 methylation status correlated with IHC (p=0.025)

HCC, hepatocellular carcinoma; CR, complete response; PR, partial response; SD, stable disease; PFS, progression-free survival; NO, nitric oxide; MOA, mechanism of action; ASS1, argininosuccinate synthetase 1; n/a, not available; PMR, partial metabolic response; OBD, optimal biological dose; BSC, best supportive care; RECIST, Response Evaluation Criteria in Solid Tumors; IHC, immunohistochemistry. a)From diagnosis.

Table 2
Phase I/II arginine depletor combination studies in advanced cancer
Study Clinical Trials. Gov Identifier No. of patients Tumor type Biomarker Lead site
Docetaxel and ADI-PEG20 NCT01497925 39 Prostate NSCLC Retrospective ASS1 University of California at Davis, CA
Cisplatin and ADI-PEG20 NCT01665183 61 Solid tumors including metastatic melanoma ASS1-deficient ADI-PEG20 MD Anderson, TX
Doxorubicin and ADI-PEG20 NCT01948843 53 Breast cancer ASS1-deficient (>75% cells) HER2 negative MD Anderson, TX
Cisplatin, pemetrexed and ADI-PEG20 To be determined 37 NSCLC (adeno) malignant pleural mesothelioma ASS1-deficient (>50% cells) FLT-PET/CT imaging Barts, London, UK

NSCLC, non-small cell lung cancer; ASS1, argininosuccinate synthetase 1; FLT, 18F-fluorothymidine; PET/CT, positron emission tomography/computed tomography.

TOOLS
PDF Links  PDF Links
PubReader  PubReader
ePub Link  ePub Link
XML Download  XML Download
Full text via DOI  Full text via DOI
Download Citation  Download Citation
  Print
Share:      
METRICS
151
Crossref
168
Scopus
19,828
View
343
Download
Related article
Editorial Office
Korean Cancer Association
Room 1824, Gwanghwamun Officia
92 Saemunan-ro, Jongno-gu, Seoul 03186, Korea
TEL: +82-2-3276-2410   FAX: +82-2-792-1410   E-mail: journal@cancer.or.kr
About |  Browse Articles |  Current Issue |  For Authors and Reviewers
Copyright © Korean Cancer Association.                 Developed in M2PI